Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Linear instability of periodic orbits of free period Lagrangian systems

  • Minimizing closed geodesics on surfaces are linearly unstable. By starting with this classical Poincaré's instability result, in the present paper we prove a result that allows to deduce the linear instability of periodic solutions of autonomous Lagrangian systems admitting an orbit cylinder (condition which is satisfied for instance if the periodic orbit is transversally non-degenerate) in terms of the parity properties of a suitable quantity which is obtained by adding the dimension of the configuration space to a suitably defined spectral index. Such a spectral index coincides with the Morse index of the periodic orbit seen as a critical point of the free period action functional in the case the Lagrangian is Tonelli, namely fibrewise strictly convex and superlinear, and it encodes the functional and symplectic properties of the problem.

    The main result of the paper is a generalization of the celebrated Poincaré 's instability result for closed geodesics on surfaces and at the same time extends to the autonomous case several previous results which have been proved by the authors (as well as by others) in the case of non-autonomous Lagrangian systems.

    Citation: Alessandro Portaluri, Li Wu, Ran Yang. Linear instability of periodic orbits of free period Lagrangian systems[J]. Electronic Research Archive, 2022, 30(8): 2833-2859. doi: 10.3934/era.2022144

    Related Papers:

    [1] Guowei Yu . Heteroclinic orbits between static classes of time periodic Tonelli Lagrangian systems. Electronic Research Archive, 2022, 30(6): 2283-2302. doi: 10.3934/era.2022116
    [2] Jiachen Mu, Duanzhi Zhang . Multiplicity of symmetric brake orbits of asymptotically linear symmetric reversible Hamiltonian systems. Electronic Research Archive, 2022, 30(7): 2417-2427. doi: 10.3934/era.2022123
    [3] Weiyu Li, Hongyan Wang . Dynamics of a three-molecule autocatalytic Schnakenberg model with cross-diffusion: Turing patterns of spatially homogeneous Hopf bifurcating periodic solutions. Electronic Research Archive, 2023, 31(7): 4139-4154. doi: 10.3934/era.2023211
    [4] Tian Hou, Yi Wang, Xizhuang Xie . Instability and bifurcation of a cooperative system with periodic coefficients. Electronic Research Archive, 2021, 29(5): 3069-3079. doi: 10.3934/era.2021026
    [5] Anna Gołȩbiewska, Marta Kowalczyk, Sławomir Rybicki, Piotr Stefaniak . Periodic solutions to symmetric Newtonian systems in neighborhoods of orbits of equilibria. Electronic Research Archive, 2022, 30(5): 1691-1707. doi: 10.3934/era.2022085
    [6] Xiaofei Zhang, Fanjing Wang . Brake orbits with minimal period estimates of first-order variant subquadratic Hamiltonian systems. Electronic Research Archive, 2022, 30(11): 4220-4231. doi: 10.3934/era.2022214
    [7] Xiaojun Huang, Zigen Song, Jian Xu . Amplitude death, oscillation death, and stable coexistence in a pair of VDP oscillators with direct–indirect coupling. Electronic Research Archive, 2023, 31(11): 6964-6981. doi: 10.3934/era.2023353
    [8] Zhili Zhang, Aying Wan, Hongyan Lin . Spatiotemporal patterns and multiple bifurcations of a reaction- diffusion model for hair follicle spacing. Electronic Research Archive, 2023, 31(4): 1922-1947. doi: 10.3934/era.2023099
    [9] Xiaochen Mao, Weijie Ding, Xiangyu Zhou, Song Wang, Xingyong Li . Complexity in time-delay networks of multiple interacting neural groups. Electronic Research Archive, 2021, 29(5): 2973-2985. doi: 10.3934/era.2021022
    [10] Minzhi Wei . Existence of traveling waves in a delayed convecting shallow water fluid model. Electronic Research Archive, 2023, 31(11): 6803-6819. doi: 10.3934/era.2023343
  • Minimizing closed geodesics on surfaces are linearly unstable. By starting with this classical Poincaré's instability result, in the present paper we prove a result that allows to deduce the linear instability of periodic solutions of autonomous Lagrangian systems admitting an orbit cylinder (condition which is satisfied for instance if the periodic orbit is transversally non-degenerate) in terms of the parity properties of a suitable quantity which is obtained by adding the dimension of the configuration space to a suitably defined spectral index. Such a spectral index coincides with the Morse index of the periodic orbit seen as a critical point of the free period action functional in the case the Lagrangian is Tonelli, namely fibrewise strictly convex and superlinear, and it encodes the functional and symplectic properties of the problem.

    The main result of the paper is a generalization of the celebrated Poincaré 's instability result for closed geodesics on surfaces and at the same time extends to the autonomous case several previous results which have been proved by the authors (as well as by others) in the case of non-autonomous Lagrangian systems.



    A celebrated result proved by Poincaré at the beginning of the last century puts on evidence the relation between the (linear and exponential) instability of an orientation preserving closed geodesic as a critical point of the geodesic energy functional on the free loop space of a surface and the minimization property of such a critical point. The literature on this criterion is quite broad. We refer the interested reader to [1,2,3] and references therein.

    A quite natural question is related to understand the role played by the energy level h on the instability properties of a periodic orbit of an autonomous Lagrangian system. Some years ago, in his expository article [4], Abbondandolo studied the question of the existence of periodic orbits of the free period Lagrangian systems in terms of the Mañé critical values. It is worth noticing that the fixed energy problem for Tonelli Lagrangians as well as for the restricted class of magnetic systems has been intensively studied by many outstanding mathematicians such as Arnold, Novikov, Ginzburg, Taimanov, Contreras, Paternain, etc. by starting on the work of Arnol'd [5] in the 1960s. If on the one hand it is impossible to keep track of all the literature on the topic, some breakthrough papers in the field are [6] and [7].

    Very recently, Ureña in [8], studied the instability of closed orbits obtained by minimization, for an autonomous Lagrangian system by combining the classical principle of Jacobi-Maupertuis principle together with a nice reduction argument firstly introduced by Carathéodory. In this way, under some suitable conditions, the dynamics of a free period Lagrangian system can be seen as the dynamics of a non-autonomous and fixed period Lagrangian system lowering by 1 the degrees of freedom.

    Let (M,,g) be a smooth n-dimensional Riemannian manifold without boundary, which represents the configuration space of a Lagrangian dynamical system. Elements in the tangent bundle TM will be denoted by (q,v), with qM and vTqM. Let L:TMR be a smooth autonomous (Lagrangian) function satisfying the following assumptions

    (N1) L is non-degenerate on the fibers of TM, that is, for every (q,v)TM we have that dvvL(q,v)0 is non-degenerate as a quadratic form, where dvvL denotes the fiberwise second differential of L

    (N2) L is exactly quadratic in the velocities meaning that the function L(q,v) is a polynomial of degree at most 2 with respect to v.

    Remark 1.1. Before stating our main result, we observe that we require that the Lagrangian function is exactly quadratic in the velocity assumption (N2). The smoothness property of L are in general not sufficient for guaranteeing the twice Fréchét differentiability of the Lagrangian action functional. In fact, the functional is twice Fréchét differentiable if and only if L is exactly quadratic in the velocities. In this case, the Lagrangian action functional is actually smooth. This fact depends upon the differentiability properties of Nemitsky operators. We observe that, by using a finite dimensional approximations for the free-period action functional as developed by authors in [7], it should be possible to remove the technical condition (N2).

    On the cartesian product Λ1(M)×R+, where Λ1(M) denotes the Hilbert manifold of 1-periodic loops on M having Sobolev regularity H1, we define the free period Lagrangian action functional as

    Eh(x,T) =T10[L(x(s),x(s)T)+h]ds,

    where h is a real constant playing the role of energy. In fact, since the system is autonomous, the energy is a first integral. By a direct computation of the first variation of the action functional Eh, it follows that (x,T) is a critical point of Eh if and only if it satisfies the Euler-Lagrangian equations having fixed energy h.

    Definition 1.2. Let (x,T) be a critical point of Eh. We term (x,T) non-null if

    dvvL(x(t),x(t))x(t),x(t)g0foreveryt[0,1].

    Moreover, (x,T) is termed

    L-Positive if dvvL(x(t),x(t))x(t),x(t)g>0;

    L-Negative if dvvL(x(t),x(t))x(t),x(t)g<0.

    We observe that the notion of L-positivity (resp. L-negativity) provides a sort of generalization of the Legendre convexity (resp. concavity) condition only along the selected orbit. Recall that the classical Legendre convexity condition reads

    (L1) L is C2 strictly convex on the fibers of TM, that is, for every (q,v)TM we get dvvL(q,v)>0 as a quadratic form.

    It is worth to observe that the above conditions have been introduced for avoiding any sign changing of the quanity

    dvvL(x(t0),x(t0))x(t0),x(t0)g.

    Otherwise, this provides deep difficulties, like in the definition of the spectral index.

    Following authors, in [9,Definition 1.2], we are ready to introduce the notion of orbit cylinder.

    Definition 1.3. A critical point (x,T) of Eh admits an orbit cylinder if there exist ϵ>0 and a smooth (in s) family critical points {(xh+s,Th+s),s(ϵ,ϵ)} of Eh+s such that (xh,Th)=(x,T). Moreover, this orbit cylinder is called non-degenerate if T(h)0.

    Under the above notation our main result reads as follows.

    Theorem 1.4. Let (x,T) be a non-null critical point of the free-period Lagrangian action Eh admitting an orbit cylinder. If one of the following four statements holds:

    (1) x is LPositive and

    (OR) x is orientation preserving and ιspec(x,T)+n is even

    (NOR) x is orientation reversing and ιspec(x,T)+n is odd

    (2) x is LNegative and

    (OR) x is orientation preserving and ιspec(x,T)+n is odd

    (NOR) x is orientation reversing and ιspec(x,T)+n is even

    then x is linearly unstable.

    Under the classical Legendre convexity condition, it is possible to prove that the spectral index reduces to the classical Morse index. So, we get the following.

    COROLLARY 1. Let us assume that (x,T) is a critical point of free period Lagrangian action Eh admitting an orbit cylinder and we assume that (L1) holds. If one of the following two alternatives holds

    (OR) x is orientation preserving and ιMor(x,T)+n is even

    (NOR) x is orientation reversing and ιMor(x,T)+n is odd

    then x is linearly unstable.

    From a tecnical viewpoint a key step for proving our main result is based on a spectral flow formula relating the Morse index of a periodic orbit as critical point of the free and fixed time Lagrangian action functional. In [9], authors actually established such a relation under a non-degenerate assumption of the orbit cylider. As already observed, here we give a generalization of the aforementioned results provided in [9] by dropping the non-degeneracy assumption of the orbit cylinder. Such a generalization is essentially based on a quite new results constructed by the second named author in [10].

    A crucial intermediate step for proving the main result of this paper is based on a spectral flow formula relating the spectral index to a symplectic invariant known in literature as Maslov index (which is an intersection invariant constructed in the Lagrangian Grassmannian manifold of a symplectic space) that plays a crucial role in detecting the stability properties of a periodic orbit. (For an index formula, we refer the interested reader to [11,12,13,14] and references therein). Very recently, new spectral flow formulas have been established and applied for detecting bifurcation of heteroclinic and homoclinic orbits of Hamiltonian systems or bifurcation of semi-Riemannian geodesics. (Cfr. [15,16,17]).

    Notation

    For the sake of the reader, let us introduce some common notation that we'll use henceforth throughout the paper.

    (M,,g) denotes a Riemannian manifold without boundary, TM its tangent bundle and TM its cotangent bundle.

    Λ1(M) is the Hilbert manifold of loops on manifold M having Sobolev regularity H1.

    ω denotes the symplectic structure J the standard symplectic matrix such that ω(,)=J, where , denotes the standard Euclidean product.

    ιMor(x) stands for the Morse index of x, ιspec(x) for the spectral index of x, ιgeo(x) for the geometrical index of x, ι1 denotes the Maslov-type index or Conley-Zehnder index of a symplectic matrix path, ιCLM denotes the Maslov (intersection) index and finally sf denotes the spectral flow.

    P(L) denotes the set of T-periodic solutions of the Euler-Lagrange Equation, P(H) the set of T-periodic solutions of the Hamiltonian Equation.

    δij is the Kronecker symbol. IX or just I will denote the identity operator on a space X and we set for simplicity Ik:=IRk for kN. Gr() denotes the graph of its argument, Δ denotes the graph of identity matrix I.

    U is the unit circle of the complex plane.

    O(n) denotes the orthogonal group, Sp(2n,R) or just Sp(2n) denotes the 2n×2n real symplectic group.

    P denotes the linearized Poincaré map.

    BFsa denotes the set of all bounded selfadjoint Fredholm operators, σ() denotes the spectrum of the operator in its argument.

    ● We denote throughout by the symbol T (resp. T) the transpose (resp. inverse transpose) of the operator . Moreover rge(),ker() and rank() denote respectively the image, the kernel and the rank of the argument.

    Γ denotes the crossing form and n+/n denote respectively the dimensions of the positive/negative spectral spaces and finally sgn() is the signature of the quadratic form (or matrix) in its argument and it is given by sgn()=n+()n().

    In this preliminary section we fix our basic notation and we explicitly provided the computation of the first and second variation of the free period Lagrangian action functional.

    Let (M,,g) be a (not necessarily compact or connected) smooth n-dimensional Riemannian manifold without boundary, which represents the configuration space of a Lagrangian dynamical system and we denote by the Riemannian norm. Elements in the tangent bundle TM will be denoted by (q,v), with qM and vTqM. The metric ,g induces: a metric on TM, Levi-Civita connections both on M and TM and finally the isomorphisms

    T(q,v)TM=Th(q,v)TMTv(q,v)TMTqMTqM,

    for Tv(q,v)TM=kerDτ(q,v) where τ:TMM denotes the canonical tangent projection.

    Notation 2.1. We shall denote by t the covariant derivative of a vector field along a smooth curve x with respect to the metric ,g. q (resp. v) denotes the partial derivative along the horizontal part (resp. vertical part) given by the Levi-Civita connection in the splitting of TTM and We shall denote by vv,qv,qq the components of the Hessian in the splitting of TTM.

    Given a positive number TR, we denote by T the one-dimensional torus T=R/TZ. Let Λ1T(M) be the Hilbert manifold of all loops y:TM having Sobolev class H1. Setting x(t)=y(tT),t[0,1] we get that the closed curve y will be identified with the pair (x,T). The action of y on time interval [0,T] is given by

    T0[L(y(s),y(s))+h]ds=T10[L(x(t),x(t)/T)+h]dt,

    where hR is a parameter. For convenience, we denote

    Eh(x,T) =T10[L(x(t),x(t)/T)+h]dt. (2.1)

    In short-hand notation, in what follows we use Λ1(M) instead of Λ11(M). In this way, we define a one-to-one correspondente between

    T>0Λ1T(M)andΛ1(M)×R+

    which preserves the action values. Bearing in mind such a correspondence, the free period action functional (2.1) is defined on the manifold Λ1(M)×R+.

    It is well-known that the tangent space TxΛ1(M) can be identified in a natural way with the Hilbert space of 1-periodic H1 (tangent) vector fields along x, i.e.,

    H(x)={ξH1(R/Z,TM)|τξ=x}.

    It is worth noticing that under the (N2) assumption, the Lagrangian functional is of regularity class C2 (actually it is smooth). Let ,1 denote the Riemannian metric on Λ1(M) defined by

    ξ,η1 =10[tξ,tηg+ξ,ηg]dt,ξ,ηH(x).

    For (ξ,b)H(x)×R, the first variation of Eh at (x,T)Λ1(M)×R+ is given by

    dEh(x,T)[(ξ,0)]=10[TqL(x(t),x(t)/T)DdtvL(x(t),x(t)/T),ξg]dt+vL(x(t),x(t)/T),ξg|10 (2.2)

    and for (0,b)H(x)×R, it reduces to

    dEh(x,T)[(0,b)]=10[(L(x(t),x(t)/T)+hvL(x(t),x(t)/T),x(t)/Tg)b]dt. (2.3)

    By Eqs (2.2) and (2.3), up to standard regularity arguments, it follows that critical points (x,T) of Eh are 1-periodic solutions of corresponding Euler-Lagrange equation having energy h. So, (x,T) satisfies the following equations:

    {Ddt(vL(x(t),x(t)/T))=TqL(x(t),x(t)/T),t(0,1)L(x(t),x(t)/T)+hvL(x(t),x(t)/T),x(t)/Tg=0

    Now, being Eh smooth it follows that the first variation dEh(x,T) at (x,T) coincides with the Fréchét differential DEh(x,T) and the second variation of Eh at (x,T) coincides with D2Eh(x,T). We set

    ˉP(t) =vvL(x(t),x(t)/T),ˉQ(t) =qvL(x(t),x(t)/T),ˉQT(t) =vqL(x(t),x(t)/T)ˉR(t) =qqL(x(t),x(t)/T),ˉL(t)=qL(x(t),x(t)/T),κ(t)=ˉP(t)x(t),x(t)g

    so, we get

    d2Eh[(ξ,b),(η,d)]==10Ddt[1TˉP(t)tξ+ˉQ(t)ξ]+ˉQT(t)tξ+TˉR(t)ξ,ηgdt+[1TˉP(t)tξ+ˉQ(t)ξ,ηg]1t=0+10{1T2ˉP(t)x(t),tηgb1T2ˉP(t)x(t),tξgd+ˉL(t)1TˉQ(t)x(t),ξgd+ˉL(t)1TˉQT(t)x(t),ηgb+1T3κ(t)bd}dt.

    Remark 2.2. For a complete details on the first and second variations, we refer the interested reader to [18] and references therein.

    For a given critical point (x,T) of Eh, let us consider the following fixed period Lagrangian action functional:

    ETh(x) =T10[L(x(t),x(t)/T)+h]dt, (2.4)

    where xΛ1(M). Actually, ETh is the restriction of Eh to the submanifold Λ1(M)×{T}. By similar calculations, the first variation of ETh at xΛ1(M) is given by

    dETh(x)[ξ]=10[TqL(x(t),x(t)/T)DdtvL(x(t),x(t)/T),ξg]dt+vL(x(t),x(t)/T),ξg|10.

    Critical points x of ETh are 1-periodic solutions of corresponding Euler-Lagrange Equation:

    Ddt(vL(x(t),x(t)/T))=TqL(x(t),x(t)/T),t(0,1). (2.5)

    The second variation of ETh at x is given by

    d2ETh(x)[ξ,η]=10Ddt[1TˉP(t)tξ+ˉQ(t)ξ]+ˉQT(t)tξ+TˉR(t)ξ,ηgdt+[1TˉP(t)tξ+ˉQ(t)ξ,ηg]10.

    By the previous computation, we get that ξkerd2ETh(x) if and only if ξ is a H2 vector field along x which solves weakly (in the Sobolev sense) the following boundary value problem

    {Ddt(1TˉP(t)tξ+ˉQ(t)ξ)+ˉQT(t)tξ+TˉR(t)ξ=0,t(0,1)ξ(0)=ξ(1),tξ(0)=tξ(1). (2.6)

    By standard bootstrap arguments, it follows that ξ is also a classical (smooth) solution of Eq (2.6).

    Remark 2.3. It is easy to prove that x is a critical point of fixed period Lagrangian system (2.4) provided that (x,T) is a critical point of free period Lagrangian system.

    The next result provides an answer to the existence of an orbit cylinder about a T-periodic orbit x. We refer the interested reader to [19,Section 4.1,Proposition 2] for the proof.

    Proposition 2.4. Let us assume that a periodic solution x(t,E) of a Hamiltonian vector field XH on M having energy E=H(x(t,E)) and period T has exactly two Floquet multipliers equal 1. Then, there exists a unique and smooth one-parameter family x(t,E) of periodic solutions with periods T(E) close to T and lying on the energy surfaces H(x(t,E))=E for |EE| sufficiently small. Moreover, T(E) converges to T(E) for EE.

    The goal of this section is to associate at the critical point (x,T) of the free period Lagrangian action given at Eq (2.1) and to a critical point x of the fixed period Lagrangian action given at Eq (2.4) the spectral index, defined in terms of the spectral flow of a suitable path of Fredholm quadratic forms. We refer the interested reader to [20,Appendix B] and references therein for a discussion about the spectral flow.

    Given (x,T) be a critical point of Eh, for any s[0,+) we let Is:(H(x)×R)×(H(x)×R)R the bilinear form defined by

    Is[(ξ,b),(η,d)] =d2Eh(x,T)[(ξ,b),(η,d)]+sα(x,T)[(ξ,b),(η,d)]whereα(x,T)[(ξ,b),(η,d)] =10{1TˉP(t)ξ,ηg+1T3κ(t)bd}dt. (3.1)

    Notation 3.1. In short-hand notation and if no confusion can arise, we set Qh =d2Eh(x,T).

    Proposition 3.2. For any s[0,+) let Qs denote the quadratic form associated to Is defined at Eq (3.1). Then

    sQs is a smooth path of Fredholm quadratic forms onto H(x)×R. In particular Qh is a Fredholm quadratic form on H(x)×R.

    Proof. The proof is completely analogous to [20,Proposition 3.2]. We let (x,T) be a critical point of the Lagrangian action functional (2.1). Then we can define the bilinear form Is,T:H(x)×H(x)R of x as a critical point of system (2.4) in the same way as (x,T). In fact, Is,T is given by

    Is,T[ξ,η] =d2ETh(x)[ξ,η]+sαT(x)[ξ,η]whereαT(x)[ξ,η] =101TˉP(t)ξ,ηgdt.

    Set QhT =d2ETh(x). Arguing precisely as done in Proposition 3.2, it can be proved Is,T and QhT both are Fredholm quadratic forms.

    For any s[0,+), let Qs,T denote the quadratic form associated to Is,T. Now we are entitled to define the following spectral indexes.

    Definition 3.3. Let (x,T) be a non-null critical point of the free period Lagrangian action given at Eq (2.3). We term spectral indices of (x,T) and x are respectively the integers given by

    ιspec(x,T) =sf(Qs,s[0,s0])andιTspec(x) =sf(Qs,T,s[0,s0]).

    where the (RHS) denotes the spectral flow of the path of Fredholm quadratic forms defined on the interval [0,s0] for a sufficiently arge s0>0.

    Remark 3.4. It is easy to show that Definition 3.3 is well-posed meaning that the spectral indices are independent on s0. This fact will be proved in the sequel and it is actually a direct consequence of Lemma 3.6.

    Proposition 3.5. We assume that (L1) holds. Then the Morse indices of (x,T) and x (i.e., the dimension of the maximal negative subspace of the Hessian of Qh and QhT) are both finite and the following equalities hold

    ιspec(x,T)=ιMor(x,T)andιTspec(x)=ιMor(x)

    Proof. We only consider the free period Lagrangian action, being the fixed period case, completely analogous. So, we start by observing that if L is C2-strictly convex on TM, then α(x,T) is a positive Fredholm quadratic form and hence sQs is a path of essentially positive Fredholm quadratic forms being realized by a path of compact symmetric perturbation of a positive definite Fredholm operator. In particular the Morse index of Qs is finite for every s[0,+). If s0 is sufficiently large the form Qs0 is non-degenerate and positive definite being the quadratic form associated to α(x,T) be Fredholm and positive definite. In particular, its Morse index vanishes. Since the spectral flow for path of essentially positive Fredholm quadratic forms is equal to the difference of the Morse indices at the ends (cf. [20,Appendix B]), we get that

    sf(Qs,s[0,s0])=ιMor(Q0)ιMor(Qs0)=ιMor(Q0).

    This concludes the proof.

    We denote by E the ,g-orthonormal and parallel frame pointwise given by

    E(t)={e1(t),,en(t)}.

    Given a critical point (x,T) of the free period Lagrangian action, we let ˉA:Tx(0)MTx(1)MTx(0)M the ,g-orthogonal operator defined by

    ˉAej(0)=ej(1).

    Such a frame E, induces a trivialization of the pull-back bundle x(TM) over [0,1] through the smooth curve x:[0,1]M; namely the smooth one parameter family of isomorphisms

    [0,1]tEtwhereEt:Rneiei(t)Tx(t)M t[0,1]andi=1,,naresuchthatEtei,Etejg=δijandtEtei=0 (3.3)

    here {ei}ni=1 is the canonical basis of Rn and δij denotes the Kronecker symbol.

    By Eq (3.3) we get that the pull-back by Et of the metric ,g induces the Euclidean product on Rn and moreover this pull-back is independent on t, as directly follows by the orthogonality assumption on the frame E.

    We set A =E10ˉA1E1O(n) and define

    Ad =[A00A]. (3.4)

    Let us now consider the Hilbert space

    H1A([0,1],Rn)={uH1([0,1],Rn)|u(0)=Au(1)}

    equipped with the inner product

    v,wA =10[v(s),Aw(s)+v(s),Aw(s)]ds.

    Denoting by Ψ:H(x)H1A([0,1],Rn) the map defined by Ψ(ξ)=u where u(t)=E1t(ξ(t)), it follows that Ψ is a linear isomorphism and it is easy to check that

    ξ(0)=ξ(1)E0u(0)=E1u(1)u(0)=Au(1)andtξ(0)=tξ(1)u(0)=Au(1)

    where, in the last, we used property of the frame being parallel.

    For i=1,,n and t[0,1], we let ei(t) =Etei and we denote by P(t),, Q(t), and R(t), respectively the pull-back by Et of ˉP(t),g, ˉQ(t),g and ˉR(t),g and y L the pull-back of ˉL by Et. Thus, we get

    P(t) =[pij(t)]ni,j=0,Q(t) =[qij(t)]ni,j=0,R(t) =[rij(t)]ni,j=0wherepij(t) =ˉP(t)ei(t),ej(t)g,qij(t) =ˉQ(t)ei(t),ej(t)g,rij(t) =ˉR(t)ei(t),ej(t)g.

    We observe that P and R are symmetric matrices and being ei(T)=nj=1aijej(0) we infer that

    P(0)=AP(T)AT,P(0)=AP(T)AT,Q(0)=AQ(T),R(0)=AR(T)AT. (3.5)

    Now, for every s[0,+), the push-forward by Ψ of the index forms Is on H(x)×R is given by the symmetric bilinear forms on H1A([0,1],Rn)×R defined by

    Is[(u,b),(v,d)]=10{1TP(t)u(t),v(t)+Q(t)u(t),v(t)+QT(t)u(t),v(t)+TR(t)u(t),v(t)}dt+10{1T2P(t)x(t),v(t)b1T2P(t)x(t),u(t)d+L(t)1TQ(t)x(t),u(t)d+L(t)1TQT(t)x(t),v(t)b+1T3κ(t)bd}dt+sα(x)[(ξ,b),(η,d)]whereα(x,T)[(u,b),(v,d)] =10{1TP(t)u(t),v(t)+1T3κ(t)bd}dt. (3.6)

    Denoting by qAs the quadratic form on H1A([0,1],Rn)×R associated to Is then, as direct consequence of Proposition 3.2, we get that for every s[0,+), the quadratic form qAs is Fredholm on H1A([0,1],Rn)×R. The following result is crucial in the well-posedness of the spectral index.

    Lemma 3.6. Under the above notation, there exists s0[0,+) large enough such that for every ss0, the form Is given in Eq (3.6) is non-degenerate (in the sense of bilinear forms).

    Proof. We argue by contradiction and we assume that for every s00 there exists ss0 such that Is is degenerate. Then there exists a (u,b)H1A([0,1],Rn)×R such that Is((u,b),(v,d))0 for every (v,d)H1A([0,1],Rn)×R, namely we have

    0Is((u,b),(v,d))=10{1TP(t)u(t),v(t)+Q(t)u(t),v(t)+QT(t)u(t),v(t)+TR(t)u(t),v(t)}dt+10{1T2P(t)x(t),v(t)b1T2P(t)x(t),u(t)d+L(t)1TQ(t)x(t),u(t)d+L(t)1TQT(t)x(t),v(t)b+1T3κ(t)bd}dt+s10{1TP(t)u(t),v(t)+1T3κ(t)bd}dt. (3.7)

    We let v(t) =1TP(t)u(t) and we observe that as direct consequence of Eq (3.5), the function v is admissible (meaning that v belongs to H1A). Since (x,T) is non-null, we set d =signκ(t)b for signκ(t){1,1}.

    By replacing v and d into Eq (3.7) respectively by v=Pu/T and d=signκ(t)b and dropping the argument t in each function, we get

    0Is((u,b),(v,d)=10{1TPu,1T(Pu+Pu)+Qu,1T(Pu+Pu)+QTu,1TPu+TRu,1TPu}dt+10{1T2Px,1T(Pu+Pu)b1T2Px,usignκ(t)b+L1TQx,usignκ(t)b+L1TQTx,1TPub+1T3κbsignκ(t)b}dt+s10{1TPu,1TPu+1T3κbsignκ(t)b}dt. (3.8)

    For i=1,2,3,4, we let Ci be given as follows:

    C1=1T2PP1+1TQP1+1TQTP1;C2=1TQP1PP1+RP1;C3=1T3Px+1T2PxP1+1TL1TQTx;C4=1T3PxPP1+L1TQxP1.

    and for for s sufficiently large we get

    0Is((u,b),(v,d)10[1T2Pu2C1PuPu+(sT2C2)Pu2C3PubC4Pub+s+1T3|κ|b2]dt>0.

    This concludes the proof.

    Remark 3.7. Here we would like to observe that the ASSUMPTION has been strongly used in the proof of Lemma 3.6 and it is crucial in the previous construction. However, it is interesting to understand if this condition is just technical or if it represent an obstruction to carry over this case Theorem 1.4.

    Now, for every s[0,+), the push-forward by Ψ of the index form Is on H(x) is given by the symmetric bilinear form on H1A([0,1],Rn) defined by

    Is,T[u,v] =10{1TP(t)u(t),v(t)+Q(t)u(t),v(t)+QT(t)u(t),v(t)+TR(t)u(t),v(t)}dtwhereαT(x)[u,v] =101TP(t)u(t),v(t)dt.

    By arguing precisely as before, we get that there exists a s0>0 large enough such that Is,T is non-degenerate for every s>s0.

    Proposition 3.8. Let (x,T) be a non-null critical point of Lagrangian action given at Eq (2.1). Then the spectral indexes are well-defined.

    Proof. We start observing that, sqAs is a path of Fredholm quadratic forms on H1A([0,T],Rn)×R. Moreover, by Lemma 3.6, there exists s0[0,+) such that qAs is non-degenerate for every ss0 and hence the integer sf{qAs,s[0,s0]} is well-defined.

    The conclusion follows by observing that qAs is the push-forward by Ψ of the Fredholm quadratic form Qs and by the fact that the spectral flow of a generalized family of Fredholm quadratic forms on the (trivial) Hilbert bundle [0,s0]×(H(x)×R) is independent on the trivialization. This concludes the proof.

    This subsection is to provide an abstract formula for computing the difference between the spectral indices defined above.

    Let H be a real separable Hilbert space equipped with inner product ,, WH be a dense subspace and let the inclusion map i:WH be compact. We let A be an unbounded linear operator on H having domain W and we assume that A is a Fredholm operator. Given a finite dimensional Hilbert space V, we assume that B:VH is a bounded linear operator and C:VV is a bounded self-adjoint linear operator. We denote by A:WVHV the self-adjoint operator defined by

    A(w,v)=(Aw+Bv,Bw+Cv),

    where B is the adjoint operator of B. In matrix form the operator A can be written as

    A=[ABBC].

    Lemma 3.9. Under above assumptions, we have

    m([0BBC])=m(C|kerB)+dim(ImB),

    where m denotes the Morse index.

    Proof. For the Hilbert space V, the following splitting holds

    V=ImBkerB=ImBkerB.

    By choosing a suitable basis, the matrix [0BBC] has the block form

    [00B1100000B110C11C1200C12C22],

    where B11:ImBImB and B11:ImBImB are both invertible. So, in particular

    [0BBC]issimilarto[00B1100000B11000000C22].

    Now, since

    m([0B11B110])=dim(ImB)andm(C|kerB)=m(C22|kerB)

    the thesis readily follows.

    Lemma 3.10. For s[0,1], we let A(s)=[A(1s)B(1s)B(1s)C] Then, the following spectral flow formula holds

    sf(A(s),s[0,1])=m(A(0)|W)+dim(WW)dim(WkerA(0)).

    Proof. We start to consider the splitting W=(kerA)kerA. So, A(s) can be written in the following block form

    A(s)=[A110(1s)B100(1s)B2(1s)B1(1s)B2(1s)C],

    where A11:(kerA)(kerA) is invertible and B1:V(kerA),B2:VkerA.

    For s[0,1], we let

    B(s)=[A110000(1s)B20(1s)B2(1s)[C(1s)B1A111B1]].

    First claim. The following equality holds:

    sf{A(s),s[0,1]}=sf{B(s),s[0,1]}. (3.9)

    This equality is a direct consequence of the stratum homotopy invariant property of the spectral flow. So, let's start to consider the 2-parameter family of operators pointwise defined by

    A(s,t)=[A110t(1s)B100(1s)B2t(1s)B1(1s)B2(1s)[C(1t)2(1s)B1A111B1]]

    and we observe that we have A(s,t)=K(t)A(s)K(t) for

    K(t)=[I0(1t)(1s)A111B10I000I].

    By a straightforward calculation it follows that dimkerA(0,t) and dimkerA(1,t) are both constants. By using the zero axiom of the spectral flow (namely each path is contained in a fixed stratum of the Fredholm Lagrangian Grassmannian), we get that

    sf{A(0,t),t[0,1]}=sf{A(1,t),t[0,1]}=0.

    By invoking the stratum homotopy invariant property of the spectral flow, we get that

    sf{A(s,0),s[0,1]}=sf{A(s,1)s[0,1]}

    which is precisely the equality appearing at Eq (3.9).

    Let

    C(s)=[0(1s)B2(1s)B2(1s)[C(1s)B1A111B1]],s[0,1].

    By taking into account the additivity property of the spectral flow under direct sum as well as of Eq (3.9), we get that

    sf{B(s),s[0,1]}=sf{C(s),s[0,1]}.

    Now, since C(1)=0, then we have

    sf{C(s),s[0,1]}=m(C(0)).

    By Lemma 3.9, we get that

    sf{C(s),s[0,1]}=m(C(0))=m((CB1A111B1)|kerB2)+dim(ImB2).

    In order to conclude, we have to prove that

    m(A(0)|W)=m((CB1A111B1)|kerB2)and
    dim(WW)dim(WkerA(0))=dim(ImB2).

    Let us consider (x1,x2,0)TkerA(0)W. Then for every (u1,u2,v)TWV we have

    [A110B100B2B1B2C][x1x20],[u1u2v]=A11x1,u1+B1x1,v+B2x2,v0. (3.10)

    We set v=0. So, A11x1,u10 for every u1(kerA) implies that A11x1=0. Consequently we have x1=0. Now, Eq (3.10) becomes B2x2,v0. Since v is arbitrary, then B2x2=0. Hence WkerA(0)={(0,x2,0)T | B2x2=0}=kerB2.

    If (x1,x2,y)TW, then for every (u1,u2,0)TW we have

    [A110B100B2B1B2C][x1x2y],[u1u20]=A11x1,u1+B1y,u1+B2y,u20. (3.11)

    We set u2=0. Then A11x1+B1y,u10 for every u1(kerA) implies that A11x1+B1y=0. Consequently we have x1=A111B1y. Let u1=0, Eq (3.11) becomes B2y,u20 for every u2kerA, then B2y=0. Hence W={(A111B1y,x2,y)T | B2y=0,x2kerA} and WW={(0,x2,0)T}=kerA.

    Now, we get

    dimWWdimWkerA(0)=dimkerAdimkerB2=dimImB2.

    For every ξ0=(A111B1y,x2,y)TW, we have

    A(0)ξ0,ξ0=[A110B100B2B1B2C][A111B1yx2y],[A111B1yx2y]=B1y,A111B1y+B1y,A111B1y+B2y,x2+B1A111B1y,y+B2x2,y+Cy,y=(CB1A111B1)y,y.

    Therefore, we have m(A(0)|W)=m((CB1A111B1)|kerB2). This concludes the proof.

    Lemma 3.11. We let A(s)=[AsBsBsC], for s[0,1] and we assume that A is invertible. Then we have

    sf(A(s),s[0,1])=m(CBA1B).

    Proof. By a similar discussion as provided in the proof of Eq (3.9), we get

    sf(A(s),s[0,1])=sf(B(s),s[0,1]), where B(s)=[A00s[CsBA1B]].

    Since A is invertible, we infer that

    sf(A(s),s[0,1])=sf(s(CsBA1B),s[0,1]).

    We observe that the operator s(CsBA1B) is defined on a finite dimensional space V, and since in this case the spectral flow is equal to the Morse index at the starting point minus the Morse index at the end point, we get that

    sf(s(CsBA1B),s[0,1])=m(CBA1B).

    By using the previous results, we are now ready to compute the difference between two spectral indices defined in Definition 3.3. By taking into account Eq (3.6), we denote by A(s) the realization of Is meaning the bounded linear operator representing the bilinear form Is w.r.t. the H1A×R scalar product; so, we get

    Is[(u,b),(v,d)]=A(s)[ub],[vd].

    Similarly, we define the bounded linear operators A(s),B,C(s) representing w.r.t. the H1A×R scalar product the three terms defining Is. So, we get

    A(s)u,v=10{1TP(t)u(t),v(t)+Q(t)u(t),v(t)+QT(t)u(t),v(t)+TR(t)u(t),v(t)}dt+s10{1TP(t)u(t),v(t)}dtB[ub],[vd]=10{1T2P(t)x(t),v(t)b+L(t)1TQT(t)x(t),v(t)b}dtC(s)b,d=(s+1)101T3κ(t)bddt.

    In matrix form the operator A(s) can be written as

    A(s)=[A(s)BBC(s)].

    Let now consider the homotopy pointwise defined by

    A(s,ϵ)=[A(s)(1ϵ)B(1ϵ)B(1ϵ)C(s)]forϵ[0,1].

    Recall the discussions below Remark 3.7 we proved that the index form Is,T is non-degenerate for s0 large enough. Therefore, A(s0) is invertible. The next result provides a more strinking propery about the spectrum of A(s0).

    Lemma 3.12. There exists δ>0 such that σ(A(s0))[δ,δ]=. In particular, the operator A1(s0) is bounded.

    Proof. Arguing by contradiction, we assume that for every δ>0 there exists λδ[δ,δ] and uδH1A([0,1],Rn) such that A(s0)uδ=λδuδ. Take vδ=Puδ, then we have

    Is0,T(uδ,vδ)=A(s0)uδ,vδ=λδuδ,vδλδ10P1Puδ2dt. (3.13)

    By taking into account Eq (3.8), we have

    Is0,T(uδ,vδ)10[1T2Puδ2C1PuδPuδ+(s0T2C2)Puδ2]dt. (3.14)

    Inequalities provided at Eq (3.13) and Eq (3.14) contradict each other for s0 sufficiently large and δ (consequently λδ) sufficiently small. This concludes the proof.

    By the homotopy invariance property of the spectral flow, we get that

    sf(A(s,0),s[0,s0])=sf(A(0,ϵ),ϵ[0,1])+sf(A(s,1),s[0,s0])+sf(A(s0,1ϵ),ϵ[0,1]). (3.15)

    Let us now compute the spectral flow sf(A(s0,1ϵ),ϵ[0,1]). By using Lemma 3.12, we infer that BA1(s0)B is a bounded operator (on a one-dimensional space). So, we get

    m(C(s0)BA1(s0)B)={1ifκ<00ifκ>0.

    By Lemma 3.11, we have

    sf(A(s0,1ϵ),ϵ[0,1])=m(C(s0)BA1(s0)B)={1ifκ<00ifκ>0. (3.16)

    Let us now compute the spectral flow sf(A(0,ϵ),ϵ[0,1]). By using Lemma 3.10 we have

    sf(A(0,ϵ),ϵ[0,1])=m(A(0,0)|W)+dim(WW)dim(WkerA(0,0))

    where W=H1A([0,1],Rn) and V=R.

    The next step is to provide an explicit description of

    m(A(0,0)|W)+dim(WW)dim(WkerA(0,0)).

    The basic idea comes from [9,Section 2.1].

    Let (x,T) be a non-null critical point of Eh with orbit cylinder (xh+s,Th+s). Then for every (ξ,b)H(x)×R we have

    dEh+s(xh+s,Th+s)[(ξ,b)]0. (3.17)

    By differentiating w.r.t. s both sides of Eq (3.17), we get

    d2Eh(x,T)[(ξh,T(h)),(ξ,b)]+s|s=0dEh+s(x,T)[(ξ,b)]=0, (3.18)

    where ξh(t)=s|s=0xh+s(t),T(h)=dds|s=0Th+s. Let now choose a variation {(xh,r,Th,r),r(ϵ,ϵ)} such that (xh,0,Th,0)=(x,T) and r|r=0(xh,r,Th,r)=(ξ,b), then we have

    s|s=0dEh+s(x,T)[(ξ,b)]=s|s=0r|r=0Eh+s(x,T)[(xh,r,Th,r)]=r|r=0s|s=0Eh+s(x,T)[(xh,r,Th,r)]=r|r=0Th,r=b.

    By taking into account Eq (3.18) we have

    d2Eh(x,T)[(ξh,T(h)),(ξ,b)]=b (3.19)

    for every (ξ,b)H(x)×R. Taking b=0 and (ξ,b)=(ξh,T(h)) respectively, we have

    d2Eh(x,T)[(ξh,T(h)),(ξ,0)]=0,d2Eh(x,T)[(ξh,T(h)),(ξh,T(h))]=T(h). (3.20)

    Let us identify H(x) with H(x)×{0} and we denote the Hessian of Eh by 2Eh(x,T). So, kerd2Eh(x,T)=ker2Eh(x,T). We now set

    H(x)={(ξ,b)H(x)×R | d2Eh(x,T)[(ξ,b),(η,0)]=0, (η,0)H(x)}.

    The following result holds.

    Lemma 3.13. Under above notations, we get

    kerd2Eh(x,T)H(x),andH(x)=kerd2Eh(x,T)R(ξh,T(h)).

    Proof. We argue by contradiction. If H(x)+kerd2Eh(x,T)=H(x)×R, then

    H(x)=(H(x)+kerd2Eh(x,T))=(H(x)×R)=kerd2Eh(x,T). (3.21)

    By taking into account Eq (3.20) we get (ξh,T(h))H(x) and by using Eq (3.19) we know (ξh,T(h))kerd2Eh(x,T) which contradicts Eq (3.21).

    So, H(x)+kerd2Eh(x,T)H(x)×R. Since dim((H(x)×R)/H(x))=1, then we finally get kerd2Eh(x,T)H(x).

    Now observe that kerd2Eh(x,T)R(ξh,T(h))H(x) and dim(H(x)/kerd2Eh(x,T))1. In particular H(x)=kerd2Eh(x,T)R(ξh,T(h)). This concludes the proof.

    By invoking Lemma 3.13, we get

    H(x)=kerd2Eh(x,T)R(ξh,T(h))H(x)H(x)kerd2Eh(x,T).

    If T(h)0, the R(ξh,T(h))H(x). Thus, we have H(x)H(x)=kerd2Eh(x,T).

    At the same time, by Lemma 3.13 we have d2Eh(x,T)|H(x)=d2Eh(x,T)|R(ξh,T(h)). Then by Eq (3.20), we have

    m(d2Eh(x,T)|H(x))={1ifT(h)>00ifT(h)<0.

    Hence, we have

    m(d2Eh(x,T)|H(x))+dim(H(x)H(x))dim(H(x)kerd2Eh(x,T))={1ifT(h)>00ifT(h)<0.

    If T(h)=0, then by Eq (3.20), we get that (ξh,T(h))H(x)H(x). Therefore, we have

    H(x)H(x)=kerd2Eh(x,T)R(ξh,0)=H(x).

    As a result, we have dim(H(x)H(x))dim(H(x)kerd2Eh(x,T))=1, and

    d2Eh(x,T)|H(x)=d2Eh(x,T)|H(x)H(x)=0.

    So, we get that if T(h)=0 there holds

    m(d2Eh(x,T)|H(x))+dim(H(x)H(x))dim(H(x)kerd2Eh(x,T))=1.

    Summing up, we have

    m(d2Eh(x,T)|H(x))+dim(H(x)H(x))dim(H(x)kerd2Eh(x,T))={1ifT(h)00ifT(h)<0.

    In conclusion, we get

    sf(A(0,ϵ),ϵ[0,1])=m(A(0,0)|W)+dim(WW)dim(WkerA(0,0))=m(d2Eh(x,T)|H(x))+dim(H(x)H(x))dim(H(x)kerd2Eh(x,T))={1ifT(h)00ifT(h)<0, (3.22)

    where W=H1A([0,1],Rn). By summarizing all the previous results, the following theorem holds.

    Theorem 3.14. Under above notations the following equalities hold:

    ιspec(x,T)ιTspec(x)=sf(Qs,s[0,s0])sf(Qs,T,s[0,s0])=sf(A(s,0),s[0,s0])sf(A(s,1),s[0,s0])=sf(A(0,ϵ),ϵ[0,1])+sf(A(s0,1ϵ),ϵ[0,1])={2ifκ<0, T(h)01ifκ<0, T(h)<0orκ>0, T(h)00ifκ>0, T(h)<0 (3.23)

    Proof. The proof readily follows by invoking Eqs (3.15)-(3.16) and Eq (3.22). This concludes the proof.

    Remark 3.15. We observe that the main role of orbit cylinder is to ensure the existence of vector (ξh,T(h)) in Eq (3.19). A natural problem is to find out some more general conditions to insure the existence of a vector in H(x) but not in kerd2Eh(x,T). The bifurcation theory of Hamiltonian system could be the right direction for answering this question.

    By using Proposition 3.5, the following result holds.

    Corollary 3.16. If L is C2-strictly convex on TM, then the difference between two Morse indices is given by

    ιMor(x,T)ιTMor(x)={1ifT(h)00ifT(h)<0

    Proof. Since L is C2-strictly convex, then κ=Px,x>0. By Proposition 3.5 and Eq (3.23), we conclude the proof.

    Remark 3.17 Here we point out that if (x,T) is a minimizer of system (2.4) when L is C2-strictly convex, then we must have ιMor(x,T)=0. By Corollary 3.16 there must hold T(h)<0. Or vice versa, if T(h)0, then (x,T) cannot be a minimizer.

    In this section we recall some well-known results about the fixed period problem. We refer the interested reader to [20] for the complete proofs.

    It is well-known that under the assumption (N1) the Legendre transform

    LL:TMTM,(q,v)(q,DL(q,v)|Tv(q,v)TM)

    is a local smooth diffeomorphism. The Fenchel transform of L is the autonomous Hamiltonian on TM

    H(q,p)=maxvTqM(p[v]L(q,v))=p[v(q,p)]L(q,v(q,p)),

    for every (q,p)TM, where the map v is a component of the fiber-preserving diffeomorphism

    L1L:TMTM,(q,p)(q,v(q,p))

    the inverse of LL.

    By the above Legendre transform, the Euler-Lagrange Equation (2.5) is changed into the following Hamiltonian equation:

    zx(t)=JH(zx(t)).

    By trivializing the pull-back bundle x(TM) over TM through the frame E defined in Subsection 3, Eq (2.6) is changed into

    {ddt(1TP(t)u(t)+Q(t)u(t))+QT(t)u(t)+TR(t)u(t)=0,t(0,1)u(0)=Au(1),u(0)=Au(1).

    By setting y(t)=1TP(t)u(t)+Q(t)u(t) and z(t)=(y(t),u(t))T we finally get

    {z(t)=JB(t)z(t),t[0,1]z(0)=Adz(1)whereB(t) =[TP1(t)TP1(t)Q(t)TQ(t)P1(t)TQT(t)P1(t)Q(t)TR(t)] (4.1)

    and Ad has been defined in Eq (3.4).

    In the standard symplectic space (R2n,ω), we denote by J the standard symplectic matrix defined by J=[0II0]. Thus the symplectic form ω can be represented with respect to the Euclidean product , by J as follows ω(z1,z2)=Jz1,z2 for every z1,z2R2n.

    Given MSp(2n,R), we denote by Gr(M)={(x,Mx)|xR2n} its graph and we recall that Gr(M) is a Lagrangian subspace of the symplectic space (R2n×R2n,ω×ω).

    Definition 4.1. Let x be a 1-periodic solution of Eq (2.5), zx be the solution of corresponding Hamiltonian equation and let us consider the path

    γΦ:[0,1]Sp(2n,R)givenbyγΦ(t) =Ad[ΦE(t)]1DϕtH(zx(0))ΦE(0).

    We define the geometrical index of x as follows

    ιgeo(x) =ιCLM(Δ,Gr(γΦ(t)),t[0,1])

    where the (RHS) in Eq (4.3) denotes the ιCLM Maslov index between the Lagrangian path tGr(γΦ(t)) and the Lagrangian path Δ =Gr(I) defined at Appendix A and references therein.

    Let x be a 1-periodic solution of Eq (2.5) and zx be the solution of corresponding Hamiltonian equation, we can define the linearized Poincaré map of zx as follows.

    Pzx:Tx(0)MTx(0)MTx(0)MTx(0)MisgivenbyPzx(α0,δ0) =ˉAd(ζ(T),1TˉP(T)tζ(T)+ˉQ(T)ζ(T))TforˉAd =[ˉA00ˉA] (4.2)

    where ζ is the unique vector field along x such that ζ(0)=α0 and 1TˉP(0)tζ(0)+ˉQ(0)ζ(0)=δ0. Fixed points of Pzx correspond to periodic vector fields along zx.

    By pulling back the linearized Poincaré map defined in Eq (4.2) through the unitary trivialization ΦE of zx(TTM) over [0,1] we get the map

    PE:RnRnRnRndefinedbyPE(y0,u0)=Ad(1TPu(T)+Qu(T),u(T))T

    where z(t)=(y(t),u(t)) is the unique solution of the Hamiltonian system given in Eq (4.1) such that z(0)=(y0,u0).

    Denoting by tψ(t) the fundamental solution of (linear) Hamiltonian system given in Eq (4.1), then we get the geometrical index given in Definition 4.1 reduces to

    ιgeo(x) =ιCLM(Δ,Gr(Adψ(t)),t[0,1]). (4.3)

    Moreover, the linearized Poincaré map can be given by the symplectic matrix Adψ(1).

    By choosing a suitable coordinates and trivialization we can split Adψ(t) into following form:

    Adψ(t)=[100tT(h)1000Px(t)],

    where Px(t) is a path of 2(n1)×2(n1) symplectic matrices. It is referred to [9,Page 104-105].

    Definition 4.2. Under above notations, zx is termed spectrally stable if the spectrum σ(Px(1))U where UC denotes the unit circle of the complex plane. If Px(1) is also semi-simple, then zx is termed linearly stable.

    Denote γ1(t)={[10tT(h)1],t[0,1]} and γ2(t)={Px(t) | t[0,1]}, then by invoking Eq (6.3), then we get

    ιgeo(x)=ιCLM(Δ,Gr(γ1(t)),t[0,1])+ιCLM(Δ,Gr(γ2(t)),t[0,1])=ιCLM(Δ,Gr(γ2(t)),t[0,1])+{1,ifT(h)<00,ifT(h)0. (4.12)

    The, next result is well-known and relates the parity of the ιCLM-index to the linear instability of the periodic orbit.

    Lemma 4.3. The following implication holds

    ιCLM(Δ,Gr(γ2(t)),t[0,1])isoddxislinearlyunstable

    Proof. It is referred to the proof of [20,Lemma 3.15].

    In [20,Equation 4.25] we give the precise relationship between ιgeo(x) and ιTspec(x):

    ιgeo(x)=ιTspec(x)+dimker(AI).

    Remark 4.4. We conclude this section, by observing that even if not explicitly stated, all arguments provided above, also work when the periodic orbit is transversally degenerate.

    Proof of Theorem 1.4. We prove only the (contrapositive of) the first statement in Theorem 1.4, being the others completely analogous. Thus, we aim to prove that

    ifxisLpositive,orientationpreservingandlinearlystableιspec(x,T)+nisodd.

    First of all, we have

    n+ιspec(x,T)=n+(ιspec(x,T)ιTspec(x))+ιTspec(x)=(ndimker(AI))+(ιspec(x,T)ιTspec(x))+(ιTspec(x)+dimker(AI))=(ndimker(AI))+(ιspec(x,T)ιTspec(x))+ιgeo(x)=(ndimker(AI))+(ιspec(x,T)ιTspec(x))+(ιgeo(x)ιCLM(Δ,Gr(γ2(t))))+ιCLM(Δ,Gr(γ2(t))). (4.5)

    Being x orientation preserving (by assumption), then detA=1 and being A also orthogonal, then we get that

    nevendimker(AI)even

    nodddimker(AI)odd.

    So, in both cases we have ndimker(AI) is even. Since x is L-Positive, then κ(t)>0. By Equations (3.23) and (4.4), we have

    T(h)0ιspec(x,T)ιTspec(x)odd and ιgeo(x)ιCLM(Δ,Gr(γ2(t)))even

    T(h)<0ιspec(x,T)ιTspec(x)even and ιgeo(x)ιCLM(Δ,Gr(γ2(t)))odd.

    So, in both cases we have (ιspec(x,T)ιTspec(x))+(ιgeo(x)ιCLM(Δ,Gr(γ2(t)))) is odd. If x is linear stable, then by taking into account Lemma 4.3, we get that ιCLM(Δ,Gr(γ2(t))) is even. Then, by Eq (4.14), we finally get that n+ιspec(x,T) is odd. This concludes the proof.

    In this section, we will give a simple example where T(h)=0 inspired by [9,Section 5].

    Let (r,θ) denote the polar coordinate on R2. Suppose D ={(r,θ) | 0<r<4} and f(r)=12(r34r2+3r), then define L:TDR by

    L(r,θ,˙r,˙θ)=12(˙r2+r2˙θ2)f(r)˙θ.

    The energy function E:TDR is given by E(r,θ,˙r,˙θ)=12(˙r2+r2˙θ2).

    By a direct computation we have γ(t) =(r(t),θ(t)) satisfies the Euler-Lagrange equations if and only if

    ¨r=˙θ(r˙θf(r));r2˙θf(r)=constant. (5.1)

    From now on we only consider the circular orbit. Suppose

    rk(t)=ρ(k),θk(t)=a(k)t,

    where ρ(k) and a(k) are both positive constants. If γ(t) is an orbit, since the second equation of (5.1) is automatically satisfied, then we only requir

    0=¨r=a(k)(ρ(k)a(k)f(ρ(k)))ρ(k)a(k)=f(ρ(k)). (5.2)

    The energy E=12(ρ2(k)a2(k))=k, then we have

    ρ(k)a(k)=2k. (5.3)

    We note that f(r)=(3r28r+3)/2, then by Equations (5.2)-(5.3), we have 2k=(3ρ2(k)8ρ(k)+3)/2, namely,

    ρ2(k)83ρ(k)+1+232k=0.

    By solving the above equation, then we get ρ(k)=4379232k.

    Consider k(12ϵ,12+ϵ), where ϵ is a sufficiently small positive number. For k=12, it is easy to calculate that

    ρ(1/2)=ρ(1/2)=1.

    By using Eq (5.3), then we have T(k)=2πa(k)=2πρ(k)2k, then

    T(k)=2π2kρ(k)ρ(k)2k2k.

    Therefore, T(1/2)=2π(ρ(1/2)ρ(1/2))=0. By Corollary 3.16, then we get

    ιMor(x,T)ιTMor(x)=1.

    The purpose of this section is to provide the symplectic preliminaries used in the paper. In Subsection A.1, we give the definition of the Maslov index. Then we compute the Maslov index of a special symplectic path. Our basic references are [13,14,21,22].

    Given a 2n-dimensional (real) symplectic space (V,ω), a Lagrangian subspace of V is an n-dimensional subspace LV such that L=Lω where Lω denotes the symplectic orthogonal, i.e., the orthogonal with respect to the symplectic structure. We denote by Λ=Λ(V,ω) the Lagrangian Grassmannian of (V,ω), namely the set of all Lagrangian subspaces of (V,ω)

    Λ(V,ω) ={LV|L=Lω}.

    It is well-known that Λ(V,ω) is a manifold. For each L0Λ, let

    Λk(L0) ={LΛ(V,ω)|dim(LL0)=k}k=0,,n.

    Each Λk(L0) is a real compact, connected submanifold of codimension k(k+1)/2. The topological closure of Λ1(L0) is the Maslov cycle that can be also described as follows

    Σ(L0) =nk=1Λk(L0)

    The top-stratum Λ1(L0) is co-oriented meaning that it has a transverse orientation. To be more precise, for each LΛ1(L0), the path of Lagrangian subspaces (δ,δ)etJL cross Λ1(L0) transversally, and as t increases the path points to the transverse direction. Thus the Maslov cycle is two-sidedly embedded in Λ(V,ω). Based on the topological properties of the Lagrangian Grassmannian manifold, it is possible to define a fixed endpoints homotopy invariant called Maslov index.

    Definition A.1. Let L0Λ(V,ω) and let :[0,1]Λ(V,ω) be a continuous path. We define the Maslov index ιCLM as follows:

    ιCLM(L0,(t);t[a,b]) =[eεJ(t):Σ(L0)]

    where the right hand-side denotes the intersection number and 0<ε<<1.

    For further reference and properties of the Maslov index we refer the interested reader to [21] and references therein.

    For the special symplectic path γ(t)=[M11(t)0M21(t)M22(t)],t[0,T], there is a very useful formula to compute its Maslov index [23,Theorem 2.2]. Here we only give the simplified version. Let V be a subspace of C2n, define

    VI={xC2n | ω(x,y)=0 yV},WI(V)={(x,u,y,v)TC4n | (x,y)TVJ,(u,v)TV}.

    Then there holds

    μCLM(WI(V),Gr(γ(t)))=m+(M1,1(T)M2,1(T)|S(T))m+(M1,1(0)M2,1(0)|S(0))+dimS(0)dimS(T), (A.1)

    where m+ denotes the Morse positive index and S(t)={xCn | (x,M1,1x)TVI}. Please note that in our situation we take K,R in [23,Theorem 2.2] as I and V respectively.

    Take V={(x,x)T | xR}, then VI=V and Δ =Gr(I)=WI(V). Let γ(t)={[10tT01],t[0,1]} where T0 is a given real constant, then we have

    ιCLM(Δ,γ(t);t[0,1])={1,ifT0>0;0,ifT00. (A.2)

    In fact, note that in this case we have S(0)=S(T)=R and M21(0)=0,M21(T)=T0, then it is just the consequence of (A.2).

    The authors thank the anonymous referees for the excellent and substantial work to evaluate the manuscript as well as for the several suggestions proposed, leading to the improvement of the manuscript to its final form.

    A special thank Prof. Xijun Hu for providing excellent working conditions in Shandong University and for the useful and deep discussions about this project.

    Li Wu is supported by NSFC (Nos.12171281, 12071255) and "The Fundamental Research Funds of Shandong University". Ran Yang is partially supported by NSFC(N.12001098) and Doctoral research start-up fund of East China University of Technology (N.DHBK2019204).

    The authors declare there is no conflicts of interest.



    [1] V. Barutello, R. D. Jadanza, A. Portaluri, Morse index and linear stability of the Lagrangian circular orbit in a three-body-type problem via index theory, Arch. Ration. Mech. Anal., 219 (2016), 387–444. https://doi.org/10.1007/s00205-015-0898-2 doi: 10.1007/s00205-015-0898-2
    [2] V. Barutello, R. D. Jadanza, A. Portaluri, Linear instability of relative equilibria for n-body problems in the plane, J. Differ. Equ., 257 (2014), 1773–1813. https://doi.org/10.1016/j.jde.2014.05.017 doi: 10.1016/j.jde.2014.05.017
    [3] X. Hu, A. Portaluri, R. Yang, Instability of semi-Riemannian closed geodesics, Nonlinearity, 32 (2019), 4281–4316.
    [4] A. Abbondandolo, Lectures on the free period Lagrangian action functional, J. Fix. Point Theory A., 13 (2013), 397–430. https://doi.org/10.1007/s11784-013-0128-1 doi: 10.1007/s11784-013-0128-1
    [5] V. I. Arnol'd, Some remarks on flows of line elements and frames, Dokl. Akad. Nauk. SSSR, 138 (1961), 255–257. https://doi.org/10.1007/978-3-642-01742-1 doi: 10.1007/978-3-642-01742-1
    [6] G. Contreras, The Palais-Smale condition on contact type energy levels for convex Lagrangian systems, Calc. Var. Partial Dif., 27 (2006), 321–395. https://doi.org/10.1007/s00526-005-0368-z doi: 10.1007/s00526-005-0368-z
    [7] L. Asselle, M. Mazzucchelli, On Tonelli periodic orbits with low energy on surfaces, Trans. Amer. Math. Soc., 371 (2019), 3001–3048. https://doi.org/10.1090/tran/7185 doi: 10.1090/tran/7185
    [8] A. J. Ureña, Instability of closed orbits obtained by minimization, Available online: https://www.researchgate.net/publication/336319646_Instability_of_closed_orbits_obtained_by_minimization
    [9] W. J. Merry, G. P. Paternain, Index computations in Rabinowitz Floer homology, J. Fix. Point Theory A., 10 (2010), 87–111. https://doi.org/10.1007/s11784-011-0047-y doi: 10.1007/s11784-011-0047-y
    [10] W. Li, The Calculation of Maslov-type index and Hörmander index in weak symplectic Banach space, Doctoral Degree Thesis, Nankai University, September, 2016.
    [11] A. Abbondandolo, A. Portaluri, M. Schwarz, The homology of path spaces and Floer homology with conormal boundary conditions, J. Fix. Point Theory A., 4 (2008), 263–293. https://doi.org/10.1007/s11784-008-0097-y doi: 10.1007/s11784-008-0097-y
    [12] X. Hu, S. Sun, Index and stability of symmetric periodic orbits in Hamiltonian systems with application to figure-eight orbit, Comm. Math. Phys., 290 (2009), 737–777. https://doi.org/10.1007/s00220-009-0860-y doi: 10.1007/s00220-009-0860-y
    [13] M. Musso, J. Pejsachowicz, A. Portaluri, A Morse index theorem for perturbed geodesics on semi-Riemannian manifolds, Topol. Methods Nonlinear Anal., 25 (2005), 69–99. https://doi.org/10.12775/TMNA.2005.004 doi: 10.12775/TMNA.2005.004
    [14] M. Musso, J. Pejsachowicz, A. Portaluri, Morse index and bifurcation of p-geodesics on semi Riemannian manifolds, ESAIM Control Optim. Calc. Var., 13 (2007), 598–621. https://doi.org/10.1051/cocv:2007037 doi: 10.1051/cocv:2007037
    [15] X. Hu, A. Portaluri, Index theory for heteroclinic orbits of Hamiltonian systems, Calc. Var. Partial Dif., 56 (2017), 1–24. https://doi.org/10.1007/s00526-017-1259-9 doi: 10.1007/s00526-017-1259-9
    [16] X. Hu, A. Portaluri, Bifurcation of heteroclinic orbits via an index theory, Math. Z., 292 (2019), 705–723. https://doi.org/10.1007/s00209-018-2167-1 doi: 10.1007/s00209-018-2167-1
    [17] X. Hu, A. Portaluri, R. Yang, A dihedral Bott-type iteration formula and stability of symmetric periodic orbits, Calc. Var. Partial Dif., 59 (2020), 1–40. https://doi.org/10.1007/s00526-020-1709-7 doi: 10.1007/s00526-020-1709-7
    [18] L. Asselle, On the existence of Euler-Lagrange orbits satisfying the conormal boundary conditions, J. Funct. Anal., 271 (2016), 3513–3553. https://doi.org/10.1016/j.jfa.2016.08.023 doi: 10.1016/j.jfa.2016.08.023
    [19] H. Hofer, E. Zehnder, Symplectic Invariants and Hamiltonian Dynamics, Birkhäuser Verlag, 2011.
    [20] A. Portaluri, L. Wu, R. Yang, Linear instability for periodic orbits of non-autonomous Lagrangian systems, Nonlinearity, 34 (2021), 237. https://doi.org/10.1088/1361-6544/abcb0b doi: 10.1088/1361-6544/abcb0b
    [21] S. E. Cappell, R. Lee, E. Y. Miller, On the Maslov index, Comm. Pure Appl. Math., 47 (1994), 121–186. https://doi.org/10.1002/cpa.3160470202 doi: 10.1002/cpa.3160470202
    [22] P. Piccione, A. Portaluri, D. V. Tausk, Spectral flow, Maslov index and bifurcation of semi-Riemannian geodesics, Ann. Global Anal. Geom., 25 (2004), 121–149.
    [23] C. Zhu, A generalized Morse index theorem, Analysis, World Sci. Publ., Hackensack, NJ, (2006), 493–540.
  • This article has been cited by:

    1. Antonio J Ureña, Instability of closed orbits obtained by minimization* , 2022, 35, 0951-7715, 5193, 10.1088/1361-6544/ac8264
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1729) PDF downloads(68) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog